Enantioselective One-Pot Synthesis of Biaryl-Substituted Amines by

Feb 12, 2019 - Chair of Organic Chemistry I, Faculty of Chemistry, Bielefeld University , Universitätsstraße 25, 33615 Bielefeld , Germany. § Insti...
0 downloads 0 Views 2MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

pubs.acs.org/journal/ascecg

Enantioselective One-Pot Synthesis of Biaryl-Substituted Amines by Combining Palladium and Enzyme Catalysis in Deep Eutectic Solvents Juraj Paris,†,‡ Aline Telzerow,§,∥ Nicolás Ríos-Lombardía,† Kerstin Steiner,§ Helmut Schwab,§ Francisco Morís,† Harald Gröger,*,‡ and Javier González-Sabín*,† †

EntreChem, SL, Vivero Ciencias de la Salud, 33011 Oviedo, Spain Chair of Organic Chemistry I, Faculty of Chemistry, Bielefeld University, Universitätsstraße 25, 33615 Bielefeld, Germany § Institute of Molecular Biotechnology, Graz University of Technology, Petersgasse 14, 8010 Graz, Austria ∥ InnoSyn B. V., Urmonderbaan 22, 6167RD Geleen, The Netherlands

Downloaded via UNIV OF NEW ENGLAND on February 12, 2019 at 22:37:53 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: The first application of Deep Eutectic Solvents (DESs) in asymmetric bioamination of ketones has been accomplished. The amine transaminases (ATAs) turned out to be particularly stable in DES-buffer mixtures at a percentage of up to 75% (w/w) neoteric solvent. Moreover, this reaction medium was used to perform a chemoenzymatic cascade toward biaryl amines by coupling a Suzuki reaction sequentially with an enantioselective bioamination catalyzed by the recently discovered ATA from Exophiala xenobiotica (EX-ωTA). The solubilizing properties of DESs enabled the metalcatalyzed step at 200 mM loading of substrate and the subsequent biotransformation at 25 mM.

KEYWORDS: Amines, Asymmetric synthesis, Biocatalysis, Deep eutectic solvents, Palladium catalysis, Suzuki cross-coupling reaction



INTRODUCTION The biaryl moiety has emerged to a broadly used, valuable structural motif not only in the field of chiral ligands for asymmetric catalysis (as underlined by BINOL and BINAP as the presumably most prominent examples bearing such a biaryl structure)1−3 but also in the fields of natural products4 and pharmaceuticals.5,6 A commercialized product in this field is Valsartane (1) developed by Ciba-Geigy (now, Novartis), which is used as an angiotensin receptor blocker for treatment of, e.g., high blood pressure (Figure 1).6 At the same time, chiral amine structural motifs can be widely found in today’s marketed drugs.7 Indeed, the FDA database reveals that 84% of the approved small molecule drugs bear at least one nitrogen atom.8 Thus, “merging” these two structural motifs has raised interest in the field of drug development, and the related biarylsubstituted amines have already turned out to represent versatile intermediates for promising drug candidates. For example, the biaryl-amine 2 is used for the synthesis of the cathepsin C inhibitor Odanacatib (3), which was evaluated in phase III for fracture prevention in postmenopausal women with osteoporosis by Merck & Co.9 This potential for pharmaceutical applications also raised the question about efficient approaches to such chiral biarylsubstituted amine molecules. Retrosynthetically, the biaryl unit can be constructed through a palladium-catalyzed Suzuki-cross © XXXX American Chemical Society

Figure 1. Examples of biaryl-containing drugs: Valsartane (1) and Odanacatib (3).

coupling reaction,10 a process which has been successfully implemented in eco-friendly media.11−13 On the other hand, the asymmetric reductive amination enables converting Received: December 21, 2018 Revised: January 24, 2019

A

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering ketones into enantiomerically pure amines.14 The combination of these two steps represents an elegant and straightforward approach toward these target molecules. The biocatalytic reductive amination can be conducted by amine transaminases.15 Such an approach has been recently demonstrated for the enantioselective synthesis of biaryl amines by the Bornscheuer group for the first time. This was exemplified for the conversion of a halogenated acetophenone with a boronic acid in a Suzuki reaction. Subsequent conversion of the resulting biaryl ketones in the presence of an amine transaminase led to the formation of amines with up to 84% overall conversion and >99% ee when conducting the two reactions sequentially at 2 mM and 1 mM substrate concentration, respectively.16 In terms of a high overall process efficiency, the combination of these two steps within a one-pot process would be highly desirable as well as the increase of the substrate loading for achieving an improved volumetric productivity. Combinations of chemo- and biocatalysis have been identified as attractive process options in recent years. This is underlined by many successful examples.17−21 Our groups have reported a related combination of a Suzuki crosscoupling reaction and subsequent enzymatic reduction in which biaryl-substituted alcohols were formed.22−24 In these studies, Deep Eutectic Solvents (DESs)25 were used as a wellknown environmentally benign solvent class which turned out to be an attractive reaction medium. In recent times, the pharmaceutical industry has become more receptive to the use of biocatalysis for the manufacture of active pharmaceutical ingredients in a sustainable manner.26 Although the greenness of this technology is typically assumed, most biotransformations suffer from issues such as water consumption, wastewater production, or unfavorable metrics due to the poor solubility of reagents in water.27 To circumvent this drawback, organic solvents can be supplemented as cosolvents, to the expense of enzyme stability, generally limited in these media. On the other hand, a new awareness has arisen in today’s synthetic organic chemistry to replace toxic/carcinogenic petroleum based volatile organic compounds (VOCs) by new, greener, and more sustainable solvents.28,29 In this context, Abbot introduced DESs,30 which consist of 2−3 compounds from renewable sources establishing an extensive H-bond network throughout the solvent with a melting point far below those of the individual components. Compared to the related ionic liquids (ILs), DESs are cheaper, easier to make, tunable, highly biodegradable, and virtually nontoxic and do not require further purification. As illustrated by the exponential growth of literature, DES technology has been applied to a wide-ranging area of research topics such as organic synthesis,31,32 metal-catalysis33−36 and organocatalysis,37−40 energy technology,41,42 material chemistry,43 or separation processes.44 In biocatalysis, since the 2008 proofof-concept,45 many examples have showcased ad hoc protocols for biotransformations in DESs and DES-buffer mixtures.46 Interestingly, hitherto there is not any example involving ATAs. Recent advances in protein engineering have enabled the conversion of a variety of sterically demanding ketones by ATAs. For example, the rationally engineered (S)-selective ATA from V. f luvialis catalyzed an ortho-biaryl ketone to the corresponding (S)-biaryl amine.47 Bornscheuer and co-workers,16 in parallel to us,48 have recently generated ATAs from Aspergillus fumigatus (4CHI-TA) and Exophiala xenobiotica (EX-ωTA), respectively, suitable for producing meta-and para(R)-biaryl amines. Thus, we became interested in developing

an efficient chemoenzymatic one-pot process route to biarylsubstituted amines based on the combination of a Suzuki cross-coupling reaction and an enzymatic transamination in DESs as sustainable reaction media. In the following, we report exactly such a process. It also represents the first example of an application of the enzyme class of amine transaminases in DESs.



EXPERIMENTAL SECTION

Reagents. D-(+)-Glucose was purchased from VWR. D-Alanine, PLP (pyridoxal 5′-phosphate hydrate), and NAD+ were purchased from Sigma-Aldrich. The ligand TPPTS (triphenylphosphine-3,3′,3′′trisulfonic acid trisodium salt hydrate, tech. 85%) was purchased from Alfa Aesar. Palladium(II) chloride was purchased from TCI. Aryl bromides and phenylboronic acids were purchased from SigmaAldrich. Enzymes. LDH (lactate dehydrogenase) from rabbit muscle (Type II, ammonium sulfate suspension, 800−1200 U/mg protein); GDH (glucose dehydrogenase) from Bacillus Megaterium was expressed in Escherichia coli. The activity was established as 1884 U/mL.49 Exophilia xenobiotica ATA gene codon-optimized for expression in Escherichia coli was ordered from Geneart/LifeTech (Vienna, Austria). The activity determined spectrophotometrically according to the acetophenone assay was EX-wt: 16.9 U/mL and EXSTA: 23.35 U/mL. Codex Transaminase ATA Screening Kit (ATASK-000250) was purchased from Codexis. Transaminases from Chromobacterium violaceum (Cv, internal plasmid number pET20) and Arthrobacter sp. [ArR (pEG23), ArS (pEG29), and ArRmut11 (pEG90)] were overexpressed in E. coli and used as lyophilized cells. The protein content of these four transaminases was established by the Pierce’s method (mg protein/mg catalyst) according to the manufacturer’s instructions: ArS (0.37), Cv (0.31), ArRmut11 (0.38), ArR (0.19). Synthesis of Deep Eutectic Solvents. ChCl-Gly (1:2 w/w), ChCl-H2O (1:2 w/w), ChCl-Sorb (1:1 w/w), and ChCl-Urea (1:2 w/ w) were prepared by gently heating under stirring at 60−80 °C for 1 h the corresponding individual components until a clear solution was obtained. General Procedure for the Bioamination of Phenylacetone. Reactions were carried out in a 2.0 mL vial. The corresponding ATA (2.0 mg for Codexis’ enzymes or 5.0 mg for Cv, ArR, ArS, and ArRmut11) was added to 500 μL of the corresponding mixture of DES and buffer 100 mM phosphate buffer pH 7.5, containing propan2-amine (1.0 M) and the cofactor PLP (1.0 mM). Then, a solution of 6 (2.0 mg) was added and the resulting mixture was shaken at 250 rpm and 30 °C for 24 h. After this time, a 50 μL aliquot was removed to determine the degree of conversion by HPLC (see section 5.2 in the Supporting Information). The reaction mixture was finally quenched with aqueous 10 N NaOH (100 μL) and extracted with ethyl acetate (2 × 500 μL). The organic layers were separated by centrifugation (90 s, 13 000 rpm), combined, and finally dried over anhydrous Na2SO4. The enantiomeric excess of the resulting amine was determined by chiral HPLC (see section 5.3 in the Supporting Information) after conventional derivatization of the sample using acetic anhydride (2 μL/mg of substrate). General Procedure for the Biarylamine Synthesis in a OnePot Process. At first, a suspension of PdCl2 (0.086 mg; 1 mol %) and TPPTS (8.32 mg; 3 mol %) in degassed and deionized water was prepared (0.25 mL). After 30 min, the resulting catalyst solution was added to a mixture consisting of aryl bromide (0.5 mmol), arylboronic acid (0.5 mmol), sodium carbonate (132 mg, 1.25 mmol), DES (2.0 mL), and degassed and deionized water (0.25 mL). The reaction mixture was heated to 100 °C (homogeneous mixture) for 24 h. After cooling to room temperature, EX-STA lysate (11 mL, 257 U), KPi buffer 100 mM pH 7.5 (6.5 mL), PLP (1 mM), NAD+ (0.1 mM), Dglucose (57 mM), D-alanine (130 mM), LDH (3600 U, 680 μL), and GDH (1200 U, 400 mg) were added. The degree of conversion and ee were determined as described above and according to sections 5.2 and 5.3 in the Supporting Information. Then, the reaction was quenched B

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering by addition of aqueous 10 N NaOH (10 mL) to adjust the pH to 14. The mixture was then extracted with ethyl acetate (2 × 25 mL), and the organic layers were separated by centrifugation (90 s, 13 000 rpm), combined, and dried over anhydrous Na2SO4 to provide the crude product. Further filtration by flash chromatography (silica gel 60 Å, ethyl acetate) yielded the corresponding (R)-biaryl amines 5a− e.

cosolvent and its ratio, the resulting amine (R)-5a displaying >99% ee in all cases. As the next step, we sought to get more insight in the unveiled stability of ATAs in DES-buffer mixtures by extending the study to enzymes from a commercial kit (Codexis),50 and also the S-selective ΑTAs from Chromobacterium violaceum (Cv)51 and (S)-Arthrobacter (ArS)52 and the R-selective ATAs from (R)-Arthrobacter (ArR)53 and its evolved variant ArRmut11.54 For this study, phenylacetone (6) was selected as a substrate, which had been efficiently converted by these ATAs in conventional aqueous medium.55 Four choline chloride-based eutectic mixtures, namely, 1ChCl/2Gly, 1ChCl/2H2O, 1ChCl/1Sorb (Sorb = sorbitol), and 1ChCl/ 2Urea were screened at variable water content (Table 2). In a typical experiment aimed at evaluating the enzymatic performance, 6 (30 mM) was incubated in a mixture of DES and potassium phosphate buffer (KPi) 100 mM (1 mM PLP and 1 M iPrNH2) at pH 7.0, 30 °C, and 250 rpm during 24 h. An initial conclusion extracted from Table 2 is that the DES-buffer mixtures resulted in highly suitable reaction media for the ATAs at 25% or 50% (w/w) DES. On the one hand, the commercial enzymes led to very high conversions in the four tested media (Table 2, entries 1−4). Indeed, the conversion values were almost identical to those reported in buffer solution,55 and in the case of ATA-256, the conversion even increased from 57% to 90−95% in the neoteric mixtures (Table 2, entry 3). Cv and ArS were less active in 1ChCl/ 2H2O, 1ChCl/1Sorb, and 1ChCl/2Urea, while they exhibited comparable activities in 1ChCl/2Gly as in buffer solution (Table 2, entries 5 and 6).55 ArR and ArRmut11 led to good conversion in all the DES-buffer mixtures, especially in the case of ArRmut11. For the particular case of 1ChCl/2Gly, a further increase to 75% (w/w) DES proved to be harmless for ATAs with changes in the conversion rate lower than 5%. The excellent tolerance of ATAs toward DESs by ATAs is noteworthy since it is far greater than toward organic solvents.56 According to a recent study, the DES nanostructure is maintained to a remarkably high level of water (ca. 42 wt % H2O) because of solvophobic sequestration of water into nanostructured domains around cholinium cations.57 Therefore, it can be assumed that the bioamination proceeds in a choline chloride/glycerol/water deep eutectic solvent mixture. Likewise, the enantioselectivity of the bioamination catalyzed by ATA-237 was enhanced from 96% to >99% ee in the four DES-buffer mixtures (see details in SI). This feature of DESs had been previously observed in KRED-catalyzed bioreductions, including an intriguing enantioselectivity switch by tuning the ratio of the DES-buffer system.58−60 Moving back to the chiral biaryl amines and keeping in mind both the reported Suzuki cross-coupling reaction in DESs and the unveiled good tolerance of EX-STA toward these solvents, we envisaged to set up a cascade combining metal catalysis and biocatalysis in such reaction media. Accordingly, we focused on the first step of the cascade, namely, the Suzuki cross-coupling reaction. Equimolar amounts of p-Br-acetophenone (8) and phenylboronic acid (9) were reacted at 40 mM in a mixture of water and different cosolvents at room temperature. As depicted in Figure 2, the measured conversions toward 4′phenylacetophenone (4b) were very high (≥85%) at 50% of iPrOH, THF, and 1ChCl/2Gly, while the reaction did not work at the same percentage of DMSO. As stated above, the negative impact of THF and i-PrOH on the catalytic performance of the EX-STA precluded the use of these



RESULTS AND DISCUSSION In the recent report, we revisited the synthesis of biaryl alcohols by means of a Suzuki cross-coupling and subsequent bioreduction of the transiently formed ketones with KREDs.24 With regard to previous research, the use of DESs enabled us to tackle the solubility hurdles and reach concentrations of 200 mM for the coupling step and 75 mM for the subsequent bioreduction. From a synthetic goal point of view, a first key challenge was to determine if EX-ωTA is active in these biobased solvents. Accordingly, the bioamination of the biaryl ketone 4a was investigated as a benchmark reaction with the variant EX-STA (amino acid exchange T273S). This variant leads to the highest conversions of biaryl ketones.48 The biotransformation was conducted under the previously optimized reaction setup, namely, based on the use of alanine as amino donor and the LDH/GDH recycling system, and supplemented with choline chloride (ChCl)/glycerol (Gly) (1:2). For the sake of comparison, other cosolvents such as DMSO, THF, and i-PrOH were also tested (5% and 15% v/v). As deduced from Table 1, the presence of THF and i-PrOH Table 1. Effect of Cosolvent on the Conversion of the ExSta-Catalyzed Bioamination of Biaryl Ketone 4aa

entry

cosolvent (%)

conv. (%)b

ee (%)c

1 2 3 4 5 6 7 8 9 10

THF (5%) THF (15%) i-PrOH (5%) i-PrOH (15%) DMSO (5%) DMSO (15%) 1ChCl/2Gly (5%) 1ChCl/2Gly (15%) 1ChCl/2Gly (25%) 1ChCl/2Gly (50%)

8 64 99 >99 >99 95 60 99 (R) n.d.d >99 (R) n.d.d >99 (R) >99 (R) >99 (R) >99 (R) >99 (R) n.d.d

a Reaction conditions: 4a (20 mM) was dissolved in the cosolvent (variable ratio) and then KPi buffer 100 mM, pH 7.5 (1 mM PLP, 0.1 mM NAD+), EX-STA (20 U), D-alanine (130 mM), glucose (60 mM), GDH (30 U), and LDH (90 U) were added and the mixture shaken for 24 h at 250 rpm and 30 °C. bMeasured by HPLC. c Measured by chiral HPLC. dNot detected.

was detrimental for the enzyme activity, resulting in a very low conversion at 15% of cosolvent (Table 1, entries 2 and 4). Conversely, the ATA remained very active in both DMSO and 1ChCl/2Gly with almost quantitative conversions at 15% of cosolvent (Table 1, entries 5−8). An increase to 25% DES turned out to decrease the enzymatic performance (60% of conversion, Table 1, entry 9), while the activity of the ATA in DES:buffer 1:1 was negligible (Table 1, entry 10). EX-STA exhibited perfect enantioselectivity toward 4a regardless of the C

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Table 2. Effect of Different DES-Buffer Media on the Conversion of the ATA-Catalyzed Bioamination of Phenylacetone (6)a,b,c

bufferd entry

ATA

1 2 3 4 5 6 7 8

ATA-237 ATA-251 ATA-256 ATA-P1-G06 Cv ArS ArR ArRmut11

95 95 57 95 91 64 91 >99

1ChCl/2Gly

1ChCl/2H2O

1ChCl/1Sorb

1ChCl/2Urea

25% DES

50% DES

75% DES

25% DES

50% DES

25% DES

50% DES

25% DES

50% DES

98 97 95 95 94 45 55 95

93 92 95 95 90 42 60 95

93 92 95 92 85 40 72 95

90 93 91 95 5 10 50 90

90 90 92 90 30 25 80 80

93 >99 85 96 12 10 85 73

88 91 88 91 15 35 90 95

91 95 90 95 99% ee (Table 3, entries 1, 5, 7, 9). Conversely, EX-wt showed poor stability in the DES-buffer medium which resulted in very low (Table 3, entries 6, 8, 10) or even no conversion (Table 3, entry 6). Although originally conceived for an enlarged small binding pocket compared to the wild-type enzyme, the resulting variant EX-STA turned out to be a more stable enzyme in DES-buffer mixtures. In conclusion, the bioamination of ketones was demonstrated in ad hoc mixtures of Deep Eutectic Solvents and aqueous buffer. The ATAs exhibited good stability and catalytic performance at very high percentages of DES (up to 75%). Owing to the unique properties of DESs, a chemoenzymatic cascade toward enantiopure biaryl amines was efficiently established. This unprecedented enzymatic activity of ATAs is an excellent proof of concept of the practical value of biorenewable solvents for synthetic chemists. Although immense advances are being made in areas such as protein engineering, it is no less true that a much simpler technique like medium engineering can be a valuable solution for optimizing a given biotransformation.



ABBREVIATIONS BINOL, 1,1′-Bi-2-naphthol; BINAP, (2,2′-Bis(diphenylphosphino)-1,1′-binaphthyl); PdCl2, Palladium chloride; TPPTS, [Tris(3-sulfonatophenyl)phosphine hydrate, sodium salt; KPi, Potassium phosphate buffer; IL, Ionic liquid; DES, Deep eutectic solvent; ChCl, Choline chloride; Gly, Glycerol; THF, Tetrahydrofuran; DMSO, Dimethyl sulfoxide; i-PrOH, Propan-2-ol; Na2CO3, Sodium carbonate; ATA, Amine transaminase; EX-ωTA, ω-Transaminase from Exophiala xenobiotica; EX-wt, Wild type EX-ωTA; EX-STA, Variant of EX-ωTA with the amino acid exchange T273S; Cv, Chromobacterium violaceum; ArS, (S)-Arthrobacter; ArR, (R)-Arthrobacter; PLP, Pyridoxal 5′-phosphate; D-Ala, D-Alanine; LDH/GDH, Lactate dehydrogenase/glutamate dehydrogenase; NAD+, Nicotinamide adenine dinucleotide; ADH, Alcohol dehydrogenase; ee, Enantiomeric excess; C, Conversion



REFERENCES

(1) Ohkuma, T.; Kurono, N. BINAP. In Privileged Chiral Ligands and Catalysts; Zhou, Q.-L., Ed.; Wiley-VCH: Weinheim, 2011; Chapter 1, p 1. (2) Shibasaki, M.; Matsunaga, S. BINOL. In Privileged Chiral Ligands and Catalysts; Zhou, Q.-L., Ed.; Wiley-VCH: Weinheim, 2011; Chapter 8, p 295. (3) Yoon, T. P.; Jacobsen, E. N. Privileged Chiral Catalysts. Science 2003, 299 (5613), 1691−1693. (4) Bringmann, G.; Walter, R.; Weirich, R. The Directed Synthesis of Biaryl Compounds: Modern Concepts and Strategies. Angew. Chem., Int. Ed. Engl. 1990, 29 (9), 977−991. (5) Nguyen, T. Giving Atropisomers Another Chance. Chem. Eng. News 2018, 96 (33), 22−25. (6) Bühlmayer, P.; Furet, P.; Criscione, L.; de Gasparo, M.; Whitebread, S.; Schmidlin, T.; Lattmann, R.; Wood, J. Valsartan, a Potent, Orally Active Angiotensin II Antagonist Developed from the Structurally New Amino Acid Series. Bioorg. Med. Chem. Lett. 1994, 4 (1), 29−34. (7) Kleemann, A.; Engels, J.; Kutscher, B.; Reichert, D. Pharmaceutical Substances: Syntheses, Patents, Applications, 4th ed.; Thieme: Stuttgart, 2001. (8) Vitaku, E.; Smith, D. T.; Njardarson, J. T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57 (24), 10257−10274. (9) Gauthier, J. Y.; Chauret, N.; Cromlish, W.; Desmarais, S.; Duong, L. T.; Falgueyret, J.-P.; Kimmel, D. B.; Lamontagne, S.; Leger, S.; LeRiche, T.; Li, C. S.; Massé, F.; McKay, D. J.; Nicoll-Griffith, D. A.; Oballa, R. M.; Palmer, J. T.; Percival, M. D.; Riendeau, D.; Robichaud, J.; Rodan, G. A.; Rodan, S. B.; Seto, C.; Thérien, M.; Truong, V.-L.; Venuti, M. C.; Wesolowski, G.; Young, R. N.; Zamboni, R.; Black, W. C. The Discovery of Odanacatib (MK-0822), a Selective Inhibitor of Cathepsin K. Bioorg. Med. Chem. Lett. 2008, 18 (3), 923−928. (10) Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95 (7), 2457−2483. (11) Marset, X.; Khoshnood, A.; Sotorríos, L.; Gómez-Bengoa, E.; Alonso, D. A.; Ramón, D. J. Deep Eutectic Solvent Compatible Metallic Catalysts: Cationic Pyridiniophosphine Ligands in Palladium Catalyzed Cross-Coupling Reactions. ChemCatChem 2017, 9 (7), 1269−1275. (12) Dilauro, G.; Garcı ́a, S. M.; Tagarelli, D.; Vitale, P.; Perna, F. M.; Capriati, V. Ligand-Free Bioinspired Suzuki−Miyaura Coupling Reactions using Aryltrifluoroborates as Effective Partners in Deep Eutectic Solvents. ChemSusChem 2018, 11 (19), 3495−3501. (13) Hooshmand, S. E.; Heidari, B.; Sedghi, R.; Varma, R. S. Recent Advances in the Suzuki−Miyaura Cross-Coupling Reaction Using Efficient Catalysts in Eco-Friendly Media. Green Chem. 2019.

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.8b06715. Additional information obtained from this study regarding the characterization of biaryl amines (NMR spectra), inhibition studies, and analytical data (PDF)



Research Article

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (J.G.-S.). *E-mail: [email protected] (H.G.). ORCID

Kerstin Steiner: 0000-0002-9037-3972 Harald Gröger: 0000-0001-8582-2107 Javier González-Sabín: 0000-0003-3764-4750 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge generous support from the European Union’s Horizon 2020 MSCA ITN-EID program under grant agreement No 634200 (Project BIOCASCADES). The authors also thank Dr. Wolfgang Kroutil for the generous gift of the amine transaminases Cv, ArS, ArR, and ArRmut11. F

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

(34) Rodríguez-Á lvarez, M. J.; Vidal, C.; Díez, J.; García-Á lvarez, J. Introducing Deep Eutectic Solvents as Biorenewable Media for Au(I)Catalysed Cycloisomerisation of γ-Alkynoic Acids: an Unprecedented Catalytic System. Chem. Commun. 2014, 50 (85), 12927−12929. (35) Iwanow, M.; Finkelmeyer, J.; Söldner, A.; Kaiser, M.; Gärtner, T.; Sieber, V.; König, B. Preparation of Supported Palladium Catalysts using Deep Eutectic Solvents. Chem. - Eur. J. 2017, 23 (51), 12467− 12470. (36) Rodrı ́guez-Á lvarez, M. J.; Vidal, C.; Schumacher, S.; Borge, J.; García-Á lvarez, J. New Iminophosphorane-Au(I) Complexes as Efficient Catalysts for the Cycloisomerization of Alkynyl Amides under Air, Room Temperature and in Aqueous or Eutectic Mixture Solutions. Chem. - Eur. J. 2017, 23 (14), 3425−3431. (37) Guajardo, N.; Müller, C. R.; Schrebler, R.; Carlesi, C.; Domínguez de María, P. Deep Eutectic Solvents for Organocatalysis, Biotransformations, and Multistep Organocatalyst/Enzyme Combinations. ChemCatChem 2016, 8 (6), 1020−1027. (38) Martínez, R.; Berbegal, L.; Guillena, G.; Ramón, D. J. Biorenewable Enantioselective Aldol Reaction in Natural Deep Eutectic Solvents. Green Chem. 2016, 18 (6), 1724−1730. (39) Massolo, E.; Palmieri, S.; Benaglia, M.; Capriati, V.; Perna, F. M. Stereoselective Organocatalysed Reactions in Deep Eutectic Solvents: Highly Tunable and Biorenewable Reaction Media for Sustainable Organic Synthesis. Green Chem. 2016, 18 (3), 792−797. (40) Ñ íguez, D. R.; Guillena, G.; Alonso, D. A. Chiral 2Aminobenzimidazoles in Deep Eutectic Mixtures: Recyclable Organocatalysts for the Enantioselective Michael Addition of 1,3-Dicarbonyl Compounds to β-Nitroalkenes. ACS Sustainable Chem. Eng. 2017, 5 (11), 10649−10656. (41) Boldrini, C. L.; Manfredi, N.; Perna, F. M.; Trifiletti, V.; Capriati, V.; Abbotto, A. Dye-Sensitized Solar Cells that use an Aqueous Choline Chloride-Based Deep Eutectic Solvent as Effective Electrolyte Solution. Energy Technol. 2017, 5 (2), 345−353. (42) Milano, F.; Giotta, L.; Guascito, M. R.; Agostiano, A.; Sblendorio, S.; Valli, L.; Perna, F. M.; Cicco, L.; Trotta, M.; Capriati, V. Functional Enzymes in Nonaqueous Environment: The Case of Photosynthetic Reaction Centers in Deep Eutectic Solvents. ACS Sustainable Chem. Eng. 2017, 5 (9), 7768−7776. (43) Carriazo, D.; Serrano, M. C.; Gutiérrez, M. C.; Ferrer, M. L.; del Monte, F. Deep-eutectic Solvents Playing Multiple Roles in the Synthesis of Polymers and Related Materials. Chem. Soc. Rev. 2012, 41 (14), 4996−5014. (44) Tang, B.; Zhang, H.; Row, K. H. Application of Deep Eutectic Solvents in the Extraction and Separation of Target Compounds from Various Samples. J. Sep. Sci. 2015, 38 (6), 1053−1064. (45) Gorke, J. T.; Srienc, F.; Kazlauskas, R. J. Hydrolase-catalyzed Biotransformations in Deep Eutectic Solvents. Chem. Commun. 2008, 1235−1237. (46) Xu, P.; Zheng, G.-W.; Zong, M.-H.; Li, N.; Lou, W.-Y. Recent Progress on Deep Eutectic Solvents in Biocatalysis. Bioresour. Bioprocess. 2017, 4 (1), 34. (47) Dourado, D. F. A. R.; Pohle, S.; Carvalho, A. T. P.; Dheeman, D. S.; Caswell, J. M.; Skvortsov, T.; Miskelly, I.; Brown, R. T.; Quinn, D. J.; Allen, C. C. R.; Kulakov, L.; Huang, M.; Moody, T. S. Rational Design of a (S)-Selective-transaminase for Asymmetric Synthesis of (1S)-1-(1,1′-Biphenyl-2-yl)ethanamine. ACS Catal. 2016, 6 (11), 7749−7759. (48) Telzerow, A.; Paris, J.; Håkansson, M.; González-Sabín, J.; RíosLombardía, N.; Schürmann, M.; Gröger, H.; Morís, F.; Kourist, R.; Schwab, H.; Steiner, K. Amine Transaminase from Exophiala Xenobiotica − Crystal Structure and Engineering of a Fold IV Transaminase that Naturally Converts Biaryl Ketones. ACS Catal. 2019, 9, 1140−1148. (49) Smith, L. D.; Budgen, N.; Bungard, S. J.; Danson, M. J.; Hough, D. W. Purification and Characterization of Glucose Dehydrogenase from the Thermoacidophilic Archaebacterium Thermoplasma Acidophilum. Biochem. J. 1989, 261 (3), 973−977. (50) The Codex Transaminase Screening Kit (Codexis, Reedwood City, USA) contains 28 transaminases (ATASK-000250).

(14) Wang, C.; Xiao, J. Asymmetric Reductive Amination. Top. Curr. Chem. 2013, 343, 261−282. (15) Guo, F.; Berglund, P. Transaminase Biocatalysis: Optimization and Application. Green Chem. 2017, 19 (2), 333−360. (16) Dawood, A. W. H.; Bassut, J.; de Souza, R. O. M. A.; Bornscheuer, U. T. Combination of the Suzuki-Miyaura-cross Coupling Reaction with Engineered Transaminases. Chem. - Eur. J. 2018, 24 (60), 16009−16013. (17) Rudroff, F.; Mihovilovic, M. D.; Gröger, H.; Snajdrova, R.; Iding, H.; Bornscheuer, U. T. Opportunities and Challenges for Combining Chemo- and Biocatalysis. Nature Catal. 2018, 1 (1), 12− 22. (18) Gröger, H. Emerging Fields in One-pot Multi-step Synthesis with Combined Chemo- and Bio-catalysts: Sequential- and Dominotype Process Concepts as well as Compartmentation Strategies. In Modern Biocatalysis: Advances Towards Synthetic Biological Systems; Williams, G., Hall, M., Eds.; The Royal Society of Chemistry: London, 2018; Chapter 15, pp 439−472. (19) Ríos-Lombardía, N.; García-Á lvarez, J.; González-Sabín, J. Onepot Combination of Metal- and Biocatalysis in Water for the Synthesis of Chiral Molecules. Catalysts 2018, 8 (2), 75. (20) Gröger, H.; Hummel, W. Combining the “Two Worlds” of Chemocatalysis and Biocatalysis towards Multi-Step One-Pot Processes in Aqueous Media. Curr. Opin. Chem. Biol. 2014, 19, 171−179. (21) Vitale, P.; Perna, F. M.; Agrimi, G.; Pisano, I.; Mirizzi, F.; Capobianco, R. V.; Capriati, V. Whole-Cell Biocatalyst for Chemoenzymatic Total Synthesis of Rivastigmine. Catalysts 2018, 8 (2), 55. (22) Burda, E.; Hummel, W.; Gröger, H. Modular Chemoenzymatic One-Pot Syntheses in Aqueous Media: Combination of a PalladiumCatalyzed Cross-Coupling with an Asymmetric Biotransformation. Angew. Chem., Int. Ed. 2008, 47 (49), 9551−9554. (23) Borchert, S.; Burda, E.; Schatz, J.; Hummel, W.; Gröger, H. Combination of a Suzuki Cross-coupling Reaction Using a Watersoluble Palladium Catalyst with an Asymmetric Enzymatic Reduction towards a One-pot Process in Aqueous Medium at Room Temperature. J. Mol. Catal. B: Enzym. 2012, 84, 89−93. (24) Paris, J.; Ríos-Lombardía, N.; Morís, F.; Gröger, H.; GonzálezSabín, J. Novel Insights into the Combination of Metal- and Biocatalysis: Cascade One-Pot Synthesis of Enantiomerically Pure Biaryl Alcohols in Deep Eutectic Solvents. ChemCatChem 2018, 10 (19), 4417−4423. (25) Smith, E. L.; Abbott, A. P.; Ryder, K. S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114 (21), 11060− 11082. (26) Devine, P. N.; Howard, R. M.; Kumar, R.; Thompson, M. P.; Truppo, M. D.; Turner, N. J. Extending the Application of Biocatalysis to Meet the Challenges of Drug Development. Nat. Rev. Chem. 2018, 2, 409−421. (27) Domínguez de María, P.; Hollmann, F. On the (Un)greenness of Biocatalysis: Some Challenging Figures and Some Promising Options. Front. Microbiol. 2015, 6, 11257. (28) Seddon, K. R. Room-Temperature Ionic Liquids: Neoteric Solvents for Clean Catalysis. Kinet. Catal. 1996, 37 (5), 693−697. (29) Bio-Based Solvents; Jeérôme, F., Luque, R., Eds.; John Wiley & Sons Ltd.: Hoboken, NJ, 2017. (30) Abbott, A. P.; Capper, D.; Davies, D. L.; Rasheed, R. K.; Tambyrajah, V. Novel Solvent Properties of Choline Chloride/Urea Mixtures. Chem. Commun. 2003, 70−71. (31) Ruß, C.; König, B. Low Melting Mixtures in Organic Snthesis − an Alternative to Ionic Liquids? Green Chem. 2012, 14 (11), 2969− 2982. (32) Zhang, Q.; de Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep Eutectic Solvents: Syntheses, Properties and Applications. Chem. Soc. Rev. 2012, 41 (21), 7108−7146. (33) Jérôme, F.; Ferreira, M.; Bricout, H.; Menuel, S.; Monflier, E.; Tilloy, S. Low Melting Mixtures Based on β-Cyclodextrin Derivatives and N,N′-Dimethylurea as Solvents for Sustainable Catalytic Processes. Green Chem. 2014, 16 (8), 3876−3880. G

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering (51) Kaulmann, U.; Smithies, K.; Smith, M. E. B.; Hailes, H. C.; Ward, J. M. Substrate Spectrum of Omega-transaminase from Chromobacterium Violaceum DSM30191 and its Potential for Biocatalysis. Enzyme Microb. Technol. 2007, 41 (5), 628−637. (52) Pannuri, S.; Kamat, S. V.; García, A. R. M. (Cambrex North Brunswick Inc.). Patent WO 2006/063336A2, 2006. (53) Mutti, F. G.; Fuchs, C. S.; Pressnitz, D.; Sattler, J. H.; Kroutil, W. Stereoselectivity of Four (R)-Selective Transaminases for the Asymmetric Amination of Ketones. Adv. Synth. Catal. 2011, 353 (17), 3227−3233. (54) Savile, C. K.; Janey, J. M.; Mundorff, E. C.; Moore, J. C.; Tam, S.; Jarvis, W. R.; Colbeck, J. C.; Krebber, A.; Fleitz, F. J.; Brands, J.; Devine, P. N.; Huisman, G. W.; Hughes, G. J. Biocatalytic Asymmetric Synthesis of Chiral Amines from Ketones Applied to Sitagliptin Manufacture. Science 2010, 329 (5989), 305−309. (55) Liardo, E.; Ríos-Lombardía, N.; Morís, F.; Rebolledo, F.; González-Sabín, J. Hybrid Organo- and Biocatalytic Process for the Asymmetric Transformation of Alcohols into Amines in Aqueous Medium. ACS Catal. 2017, 7 (7), 4768−4774. (56) Leipold, L.; Dobrijevic, D.; Jeffries, J. W. E.; Bawn, M.; Moody, T. S.; Ward, J. M.; Hailes, H. C. The Identification and Use of Robust Transaminases from a Domestic Drain Metagenome. Green Chem. 2019, 21 (1), 75−86. (57) Hammond, O. S.; Bowron, D. T.; Edler, K. J. The Effect of Water upon Deep Eutectic Solvent Nanostructure: An Unusual Transition from Ionic Mixture to Aqueous Solution. Angew. Chem., Int. Ed. 2017, 56 (33), 9782−9785. (58) Maugeri, Z.; Domínguez de María, P. Whole-Cell Biocatalysis in Deep-Eutectic-Solvents/Aqueous Mixtures. ChemCatChem 2014, 6 (6), 1535−1537. (59) Vitale, P.; Abbinante, V.; Perna, F. M.; Salomone, A.; Cardellicchio, C.; Capriati, V. Unveiling the Hidden Performance of Whole Cells in the Asymmetric Bioreduction of Aryl-containing Ketones in Aqueous Deep Eutectic Solvents. Adv. Synth. Catal. 2017, 359 (6), 1049−1057. (60) Cicco, L.; Ríos-Lombardía, N.; Rodríguez-Á lvarez, M.; Morís, F.; Perna, F. M.; Capriati, V.; García-Á lvarez, J.; González-Sabín, J. Programming Cascade Reactions Interfacing Biocatalysis with Transition-metal Catalysis in Deep Eutectic Solvents as Biorenewable Reaction Media. Green Chem. 2018, 20 (15), 3468−3475.

H

DOI: 10.1021/acssuschemeng.8b06715 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX